Você está na página 1de 49

CONCREEP 10 211

Concrete Pavement Joint Durability: A SorptionBased Model for Saturation,


the Role of Distributed Cracking and Calcium Oxychloride Formation

W. Jason Weiss1
1
Jack and Kay Hockema Professor, School of Civil Engineering, School of Materials
Engineering, Purdue University, 550 Stadium Mall Dr., West Lafayette, IN 47906.
E-mail: wjweiss@purdue.edu

Abstract

Concrete pavement joints have shown the potential to deteriorate, thereby


compromising the service life of an otherwise healthy pavement. It is proposed that
this damage is due to fluid collecting at the joint. The collection of fluid makes
concrete 1) susceptible to freeze and thaw damage when it reaches a critical degree of
saturation and 2) susceptible to damage due to a chemical reaction between the salt
and matrix. Currently, models that describe the rate of water absorption in concrete
consider the concrete as uncracked. In service however, concrete can develop cracks
that increase the rate of transport. The first part of this paper will begin by discussing
water absorption in undamaged concrete. The second part of this paper discusses the
influence of distributed cracking. The third part of this paper discusses a potential
chemical reaction between cement and deicing salt that forms calcium oxychloride.

BACKGROUND ON THE PROBLEM OF CONCRETE JOINT DAMAGE

While many concrete pavements in North America provide excellent long-term


performance, a portion of the concrete pavements have exhibited premature
deterioration primarily at the saw-cut joints (Jones et al. 2014). Joint deterioration is
problematic because it compromises the performance and service life of an otherwise
healthy pavement. Repairing joint damage can be expensive and these repairs are
disruptive to the travelling public. It is commonly observed that the joints that
frequently exhibit this damage contain standing fluid. Two primary types of
deterioration can be observed: 1) cracking parallel to the joint resulting in 2 mm to 12
mm flakes of concrete being removed from the joint or 2) spalling that begins with a
hollowed out region at the bottom of the saw cut resulting in a crack to the surface of
pavement that is typically 75 to 150 mm from the edge of the joint. This paper
provides a short update of recent research investigating the role of 1) saturation and
freezing, 2) the influence of distributed cracking on the rate of saturation and
freezing, and 3) the reaction between deicing salts (in this case calcium chloride) and
the cementitious matrix.

© ASCE Rev. 11/2014


CONCREEP 10 212

A WATER ABSORPTION BASED MODEL FOR FREEZE THAW DAMAGE

Fagerlund (1977) pioneered the use of an approach for modeling freeze-thaw damage
that considered that concrete would not be able to withstand saturation levels above a
certain value. This value has been measured to be 86% for some concrete (Li et al.
2012) and will be assumed to be 85% for this paper. A two stage absorption approach
was used to describe the rate at which concrete reached a critical level of saturation.
Bentz et al. (2011) simplified the approach assuming a single sorption rate by
assuming a weather event (e.g., rain) of longer than a particular time (6 hrs) was
required to reach the second level of saturation. Bentz et al. (2001) used this
approach with weather data to predict freeze-thaw damage in concrete pavements.
Lucero et al. (in press) used neutron radiography to illustrate that the initial sorption
was related to the filling of gel and capillary pores while the secondary rate of
sorption was related to the filling of larger pores like air voids (Todak et al. 2015).

This paper examines the role of the mixture proportions on saturation response. It is
assumed that the transition between the initial absorption and secondary absorption
(commonly referred to as the nick point) can be described by a value of saturation
that fills in all the pores in a concrete with the exception of the air voids. Powers
model (Powers and Brownyard 1942) can be used to determine the degree of
saturation at the nick point (Figure 1a) where the degree of saturation is shown as a
function of the entrained air volume. Concrete with more air has a lower degree of
saturation (i.e., is further away from reaching critical saturation). It has been shown
that when water is in contact with a small sample of concrete (50 mm thick sample)
the nick point will generally occur during the first 24 hours for typical paving
concrete (Lucero et al. in press). This value can be used with the secondary rate of
sorption to predict the time required to reach the critical degree of saturation as
shown in equation 1:

ܵ஼ோ ൌ ܵேூ஼௄ ൅ οܵξ‫ݐ‬ Equation 1

where SCR is the critical degree of saturation, SNICK is the degree of saturation
between the initial and secondary absorption which occurs when all the pores are
filled except the air entrainment, ΔS is the secondary rate of absorption and t is time.

While SNICK is dependent on the entrained air volume, the secondary rate of
absorption is strongly related to the quality of the matrix. Figure 1b illustrates the
impact of the sorption rate on the time to achieve critical saturation for a concrete
made with an air content of 5.5% and a water to cement ratios (w/c) of 0.39 to 0.45
(with SNICK varying from 66% to 70% respectively. The range of secondary sorption
was measured by Castro et al. (2011) with 0.033 being representative of a mixture
with a w/c of 0.39 and 0.053 being representative of a mixture with a w/c of 0.45.

Figure 2a provides an illustration of the influence of air content on the time it takes to
saturate the concrete. It can be seen that the volume of air has a substantial influence

© ASCE Rev. 11/2014


CONCREEP 10 213

on the time to reach critical saturation. Figure 2b illustrates the influence of w/c
where the impact is due to the change in the rate of water absorption.

1 40
Degree of Saturation (~)

0.8
30

0.6
20
0.4

10
0.2

0 0
0 2 4 6 8 10 0.02 0.03 0.04 0.05 0.06
Volume of Air (%) Secondary Sorption Rate (%DOS/year0.5)

(a) (b)

Figure 1. a) Influence of Air Content on Degree of Saturation and b) Influence of


the Secondary Rate of Sorption on the Time to Reach Critical Saturation

1.0
Degree of Saturation (~)

0.9

0.8

0.7

0.6 4% Air
6% Air
8% Air
0.5
0 10 20 30 40
Time (Years)

(a) (b)

Figure 2. a) The Illustration of the Air Content on the Degree of Saturation and
b) the Influence of Water to Cement Ratio on the Degree of Saturation

A WATER ABSORPTION BASED MODEL THAT CONSIDERS CRACKING

While the previous section of the paper discussed the time to reach the critical degree
of saturation for a concrete that was undamaged, this portion of the paper will begin

© ASCE Rev. 11/2014


CONCREEP 10 214

to discuss the role that cracking or damage may have in accelerating the rate of water
absorption. Previous research by Yang et al. (2006, 2007) has indicated that there are
two very distinctive types of crack morphology that can occur in a concrete
pavement. Crack morphology can be divided into two main types of cracking: 1)
localized discrete cracks which are consistent with mechanical loading and 2)
distributed crack networks which are consistent of those occurring due to freeze thaw
damage or alkali silica reactivity. The localized cracking can result in localized water
ingress creating regions with higher degrees of saturation. The distributed crack
networks are influenced by the crack distribution, orientation, and morphology. The
cracked networks influence the overall rate of fluid absorption. Yang et al. (2006)
showed that a cracked network that was consistent with connected cracking and a
decrease in sample stiffness of 30% resulted in a sorption that was 1.9 times higher
than that of the undamaged concrete. This is more complicated when the crack
network is not completely connected. Figure 3a illustrates the influence of damage
on the projected time to saturation. Figure 3b illustrates the projected time for the
concrete to reach critical saturation. Time to Critical Saturation (Years)

1.0 60
Undamaged
Concrete
Degree of Saturation (~)

0.9 (E/E0 = 1.00)


Damaged
40 Concrete
0.8 (E/E0 = 0.69)

0.7
20

0.6
Damaged Concrete
Undamaged Concrete
0.5 0
0 10 20 30 40 0 2 4 6 8 10
Time (Years) Volume of Air (%)

(a) (b)

Figure 3. a) The Influence of Damage on the Degree of Saturation and


b) The Influence of Damage on the Time to Reach Critical Saturation

CHEMICAL REACTIONS BETWEEN DEICING SALT AND CONCRETE

The fluid that collects in the pavement joints is not pure water but contains deicing
salts. While the use of deicing salts is an effective method for melting the ice and
increasing the safety to the travelling public, it may also be partially responsible for
joint deterioration issues that develop at the joints in the pavement. The reactions
that occur depend on the salt that is used. It appears that NaCl results in an expansive
and damaging sulfo-aluminate phase (Farnam et al. 2014a) while CaCl2 and MgCl2
result in the formation of a calcium oxychloride (Farnam et al. 2014b, Farnam et al.
2015, Farnam et al. submitted, Sutter et al. 2008, Villani et al. 2015).

© ASCE Rev. 11/2014


CONCREEP 10 215

A procedure was developed to measure the heat flow of a salt-hydrated cement


powder sample using low-temperature differential scanning calorimetry (LT-DSC)
(Monical et al. in preparation). The LT-DSC test is performed by cooling the sample
to -80C, performing a small cooling loop, then reheating the material at a rate of 0.3
C/min while measuring the heat release. Three main events occur resulting in heat
release. At approximately -50C heat is released as the eutectic solid changes phase
becoming a highly concentrated salt liquid (the concentration is the eutectic
composition). In the range between the eutectic temperature to the liquidus
temperature a second feature in the heat flow curve occurs that is related to the
melting of ice. The third feature in the heat flow curve occurs at temperatures
between those associated with the melting of ice and approximately 50C are
associated with the formation of calcium oxychloride. By integrating the heat flow
peak associated with the calcium oxychloride the total heat released can be obtained.
Oxychloride requires the presence of calcium chloride from the salt and calcium
hydroxide from the cementitious matrix. Figure 4 compares the amount of calcium
oxychloride that forms for cement matrixes made using a variety of cementitous
binders. It can be observed that when supplementary cementitious materials are used
(slag, fly ash or silica fume) the amount of calcium oxychloride that is produced
decreases and the pavement joint would be expected to be more durable.

Figure 4. An Illustration of the Composition of the Matrix and the Potential


Formation of Calcium Oxychloride

© ASCE Rev. 11/2014


CONCREEP 10 216

DISCUSSION ON IMPROVING JOINT DURABILITY IN CONCRETE

The durability of concrete pavement joints in North America have shown signs of
deterioration. This deterioration has been previously proposed by the author (2005)
to be related to fluid collecting in the joint that can lead to hydraulic pressure during
freezing and/or a chemical reaction between the deicing salt and matrix. The fluid
collects when the joint sealant fails and the joint does not crack or the crack does not
open to a sufficient width to allow it to drain. Many agencies are searching for ways
to mitigate this damage. The increase of the air content has been proposed which
may help to reduce the initial degree of saturation and increase the time for the
concrete to reach a critical saturation. This has been shown to be an effective way to
reduce the degree of saturation when the pores in the matrix saturate and this is
beneficial. It should be noted however that the volume of the air void system (and
saturation) may be compromised when the deicing salt enters the pavement reacting
with the calcium aluminate phase resulting in Friedels salt or Kuzel’s salt. It has also
been proposed that decreasing the water – to cement ratio of the pavement can
increase life by reducing the rate of fluid absorption. While both of these approaches
are valid their benefits can not completely solve the problem. It has been shown that
damage can be caused by saw-cutting (Raoufi et al. 2008) in a heart shaped region at
the base of the saw-cut. It is proposed that this damaged region could lead to an
increase in the rate of saturation and damage. By improving the timing of saw-
cutting using maturity to minimize damage at the base of the saw cut has great
potential. Coates et al. (2008) and Golias et al (2012) have suggested that the use of
a concrete sealant can reduce the rate of fluid ingress and thereby reducing the rate of
fluid ingress and potential for the concrete to reach critical saturation or the potential
chemical reaction between the salt and concrete resulting in an increase in the
durability of the concrete pavement joint. The use of supplementary cementitious
materials (SCM, fly ash, silica fume, slag) is beneficial. While use of SCM reduces
the transport properties (similar to reducing the water to cement ratio) they can also
reduce the potential for reaction between calcium chloride and calcium hydroxide.

CONCLUSION

Concrete is susceptible to freeze and thaw damage when it reaches a critical degree of
saturation. Currently, most models that describe the rate of water absorption in
concrete consider the concrete as uncracked. In service however, concrete can
develop cracks. These cracks can dramatically influence the transport of liquids and
ions. However not all cracks are created equal. This paper will discuss the rate at
which water is absorbed in concrete and the rate at which concrete reaches this
critical degree of saturation. Specifically the first portion of this paper will begin by
discussing water absorption in undamaged concrete. In the second part of the paper
the influence of cracking was discussed on the rate of water absorption. Specifically,
crack morphology will be divided into two main sources of cracking: 1) localized
discrete cracks and 2) distributed crack networks. This paper used data that shows
how distributed microcracking, similar to the cracking that is expected during freeze
thaw or alkali silica reaction, can alter the rate of absorption. Finally the potential

© ASCE Rev. 11/2014


CONCREEP 10 217

damage that occurs due to a reaction between the deicing salt (calcium chloride) and
the calcium hydroxide in the cementitious matrix.

ACKNOWLEDGEMENTS

This work was supported in part by the Joint Transportation Research Program
administered by the Indiana Department of Transportation (SPR 3864) and the
Portland Cement Association/National Ready Mix Concrete Association through the
CSH HUB initiative. The contents of this paper reflect the views of the authors, who
are responsible for the facts and the accuracy of the data presented herein, and do not
necessarily reflect the official views or policies of the Federal Highway
Administration, the Indiana Department of Transportation, Portland Cement
Association or National Ready Mix Concrete Association, nor do the contents
constitute a standard, specification, or regulation.

REFERENCES

Bentz, D. P., Ehlen, M., Ferraris, C., Garboczi, E., (2001) “Sorptivity-Based Service
Life Predictions for Concrete Pavements,” International Conference on
Concrete Pavements, September 2001
Castro, J., Bentz, D., & Weiss, J. (2011). Effect of sample conditioning on the water
absorption of concrete. Cement and Concrete Composites, 33(8), 805–813.
doi:10.1016/j.cemconcomp.2011.05.007
Coates, K., Mohtar, S., Tao, B., and Weiss, W. J., (2009) “Can Soy Methyl Esters
Reduce Fluid Transport and Improve the Durability of Concrete?”
Transportation Research Board, Volume 2113, pp. 22-30
Fagerlund, G. (1977). The international cooperative test of the critical degree of
saturation method of assessing the freeze / thaw. Materials and Structures,
10(4), 231–253.
Farnam, Y., Bentz, D. P., Hampton, A., & Weiss, W. J. (2014). Acoustic Emission and
Low Temperature Calorimetry Study of Freeze and Thaw Behavior in
Cementitious Materials Exposed to Sodium Chloride Salt. Journal of the
Transportation Research Record, 2441, 81–90.
Farnam. Y., Bentz, D. P., Sakulich, A., Flynn, D., and Weiss, W. J., (2014) “Using A
Low Temperature Guarded Comparative Longitudinal Calorimeter and
Acoustic Emission to Measure Freeze and Thaw Damage in Mortars
Containing Deicing Salt” Journal of Advances in Civil Engineering Materials,
ASTM, Vol. 3, No. 1, 2014, pp. 316–337, doi:10.1520/ACEM20130095
Farnam, Y., Wiese, A., Bentz, D., Davis, J. and Weiss, W. J., (2015), Damage
Development in Cementitious Materials Exposed to Magnesium Chloride
Deicing Salt, Accepted for Publication in the Journal of Construction and
Building Materials, Elsevier (In Press).
Farnam, Y., Dick, S., Wiese, A., Davis, J., Bentz, D. P. and Weiss, W. J., (Submitted)
“Influence of Calcium Chloride Deicing Salt on Phase Changes and Damage
Development in Cementitous Materials,” Cement & Concrete Composites

© ASCE Rev. 11/2014


CONCREEP 10 218

Farnam, Y. Esmaeeli, H.S. Bentz, D. Zavattieri, P. and Weiss J. (2015) “Experimental


and Numerical Investigation on the Effect of Cooling/Heating Rate on the
Freeze-Thaw Behavior of Mortar Containing Deicing Salt Solution,”
International Conference on the Regeneration and Conservation of Concrete
Structures (RCCS), Nagasaki, Japan, June 2015, pp. 1-12
Golias, M., Castro, J., Peled, A., Nantung, T., Tao, B., and Weiss, J., (2012) “Can Soy
Methyl Ester (SME) Improve Concrete Pavement Joint Durability,”
Transportation Research Record, Vol. 10 p. 60-68
Jones, W., Y. Farnam, P. Imbrock, J. Spiro, C. Villani, M. Golias, J. Olek, and W. J.
Weiss. An Overview of Joint Deterioration in Concrete Pavement:
Mechanisms, Solution Properties, and Sealers. Purdue University, West
Lafayette, Indiana, 2013. doi: 10.5703/1288284315339
Li, W., Pour-Ghaz, M., Castro, J., & Weiss, J. (2012). Water Absorption and Critical
Degree of Saturation Relating to Freeze-Thaw Damage in Concrete Pavement
Joints. Journal of Materials in Civil Engineering, (March), 299–307.
doi:10.1061/(ASCE)MT.1943-5533.0000383.
Lucero, C., Bentz, D. P., Hussey. D., Jacobsen, D. and Weiss, W. J. (in press) “Using
Neutron Radiography to Quantify Water Transport and the Degree of
Saturation in Entrained Air Cement Based Mortar,” Physics procedia
Monical, J., Villani, C., Farnam, Y., Unal, E., and Weiss, W. J., (in preparation)
“Using Low Temperature Differential Scanning Calorimetery to Quantify
Calcium Oxychloride Formation for Different Cementitious Materials in the
Presence of CaCl2,” ASTM Advances in Civil Engineering Materials
Raoufi, K., Radlinska, A., Nantung, T., and Weiss, W. J., (2008) “Practical
Considerations To Determine The Time And Depth Of Saw-Cuts In Concrete
Pavements,” Transportation Research Record
Spragg, R., Castro, J., Li, W., Pour-Ghaz, M., Huang, P., and Weiss, W. J., (2011)
“Wetting and Drying of Concrete in the Presence of Deicing Salt Solutions”,
Cement and Concrete Composites, Volume 33, Issue 5, May, Pages 535-542
Sutter, L., Peterson, K., Julio-Betancourt, G., Hooton, D., Van Dam, T., & Smith, K.
(2008). The deleterious chemical effects of concentrated deicing solutions on
Portland cement concrete. Pierre, South Dakota
Todak, H., Lucero, C., and Weiss, W. J., (2015) “Why is the Air There? Thinking
about Freeze-Thaw in Terms of Saturation,” Concrete In Focus
Villani, C., Farnam, Y., Washington, T., Jain, J., & Weiss, W. J. (2015). Performance
of Conventional Portland Cement and Calcium Silicate Based Carbonated
Cementitious Systems During Freezing and Thawing in the presence of
Calcium Chloride Deicing Salts. Transportation Research Record
Yang, Z., Weiss, J., Olek, J., & Lafayette, W. (2007). Water absorption in partially
saturated fractured concrete. In RILEM workshop: transport mechanism in
cracked concrete (pp. 1–8). Ghent
Yang, Z., Weiss, W. J., and Olek, J., (2006) “Water Transport in Concrete Damaged
by Tensile Loading and Freeze-Thaw Cycling,” ASCE Journal of Civil
Engineering Materials, Vol. 18, No. 3, pp. 424-434

© ASCE Rev. 11/2014


CONCREEP 10 219

Prediction of the Time-Variant Behaviour of Concrete Sewer Collection Pipes


Undergoing Deterioration 'ue to Biogenic Sulfuric Acid

L. ěoutil1; M. Chromá1; B.Teplý1; and D. Novák1


1
Faculty of Civil Engineering, Brno University of Technology, VeveĜí 95, 60200
Brno, Czech Republic. E-mail: routil.l@fce.vutbr.cz; chroma.m@fce.vutbr.cz;
teply.b@fcevutbr.cz; novak.d@fce.vutbr.cz

Abstract

The influence of microbiologically induced corrosion on the bearing capacity


of concrete sewer pipes is modeled as a time dependant and stochastic problem. The
potential of the model and software is shown in this presentation of a parametric
study of a corroding pipe, the comparison of crushing strength to test results and an
approach for the complex analysis of a buried pipe.

INTRODUCTION

In sewage collection systems a corrosion problem exists involving the


destruction of concrete pipes or structures by sulfuric acid produced by sulfur-
oxidizing bacteria from hydrogen sulfide. This type of corrosion is termed
microbiologically induced corrosion (MIC), or more precisely biogenic sulfuric acid
corrosion (Neville 2004), (Yuan et al. 2013), and its effect may be intensified in
combination with mechanical loading actions (RILEM 2013). The influence of this
corrosion on the service life, bearing and deformation capacity of such concrete
structures is obvious. Tremendous resources are being spent worldwide on the
maintenance and repair of sewer networks. The degradation problems affecting
concrete sewer pipes and the modelling of such problems were previously dealt with
e.g. in (Tee et al. 2011), (Yamamoto et al. 2013), (Wells & Melchers 2014) and (De
Belie et al. 2004).
In the context of concrete degradation processes, time is the decisive variable
and the durability issues affecting concrete structures are significant. With respect to
the tremendous number of possible variants and combinations as regards materials,
exposure, loading and structure type, it is not feasible to perform a comprehensive
investigation using experimental techniques. Analytical modelling is a reasonable
alternative which also enables the consideration of the existing scatter of different
quantities involved. This is also clearly reflected in recent standardization activities
where design for durability is dealt with: (ISO 13823 2008), (fib 2012) and (ISO
16204 2012). These documents deal with probabilistic approaches and introduce the
design of structures for durability – i.e. a time-dependent limit state approach which
takes service life into account. It appears that predictive models are needed to
estimate how resistance (and/or loads) will change over time, and the involved
uncertainties (both epistemic and eleatoric) need to be given proper consideration.

© ASCE Rev. 11/2014


CONCREEP 10 220

In (Tee et al., 2011) a relatively detailed stochastic analysis of corrosion-


affected concrete sewer pipes is presented and the expected service life of pipes
subjected to hydrogen sulfide-induced corrosion is predicted. However, structural
failures are determined in a rather simplified way only. In (Yamamoto et al. 2013) the
method of estimating bending strength based on endoscopic camera diagnostics and
Japanese design recommendations is discussed; simplified formulae for structural
analysis are employed and only a deterministic formulation is provided.
The present paper aims at presenting a complex stochastic method for the
assessment of bearing capacity (in the sense of the serviceability, ultimate effects and
durability) of concrete sewer pipes. The modeling approach considers the effects of
the combination of biogenic sulfuric acid corrosion with mechanical loading actions
and its influence on the structural performance of concrete sewer pipes. Structural
changes over time due to the MIC corrosion of a circular structure are studied.

BIOGENIC SULFURIC ACID CORROSION

Large amounts of concrete and cementitious materials are used in wastewater


systems and sulfate attack is one of the major threats to the durability of such
infrastructure. Concrete pipes, walls and other structures can be affected by sulfate
directly from wastewater (in the bottom part of pipelines) or produced by
microbiological activity (on the sewer crown). Although there are various
microorganisms in sewage pipes which can generate many kinds of aggressive
species which may damage concrete and reduce the material’s service life, the most
prominent biodeterioration is caused by biogenic sulfuric acid corrosion (Neville
2004), (Yuan et al. 2013), (Belie et al. 2004). Chemical sulfate attack on concrete
results in the neutralisation of the concrete surface, the cracking and scaling of
concrete material accelerated by the sewage flow, and the disruption of the concrete
and the significant loss of its mechanical strength (Sun et al. 2013), (Sarkar et al.
2010), (Stein 1999). From this aspect, sulfate attack in a sewage system is very
dangerous, especially in areas where waste water is rich in sulfates or H2S.

LIMIT STATES, MODELLING AND A SOFTWARE TOOL

Limit states
The Ultimate Limit State (ULS) and the Serviceability Limit State (SLS) are
generally evaluated while designing/assessing a concrete structure. The general
condition for the probability of failure Pf reads:
ܲ௙ ൌ ܲሺܵ ൒ ܴሻ ൏ ܲௗ (1)
where S is the action effect, R is the resistance (barrier) and Pd is the design
(acceptable, target) probability value. The index of reliability is alternatively utilized
instead of Pf in practice.
Generally, both S and R (and hence Pf) are time dependent and random.
Durability is related to the design working life tD (or service life) (Teplý &
VoĜechovská 2012). Two types of Pf assessment can be distinguished that consider the
randomness of concrete degradation due to biogenic sulfuric acid corrosion:

© ASCE Rev. 11/2014


CONCREEP 10 221

(i) Focusing on the pipe wall only, in which case the following holds for the
action effect (for a degraded concrete layer)
S=ct (2)
where c stands for the rate of corrosion; R in condition (1) represents in this case for a
value limiting the thickness of the concrete pipe wall in the sense of ULS (bearing
capacity) or SLS (deflection limit or crack opening). See Section “Parametric study”
for an example.
(ii) In a more complex case the action effect S is the bearing capacity of the
(buried) pipe (ULS or SLS) with a degraded wall, under given conditions and load;
the barrier R is a relevant limiting value.

Modeling of concrete degradation


The rate of concrete deterioration depends, among other things, upon local
conditions and material characteristics. The degradation of concrete sewer pipes due
to sulfuric acid attack can be assessed by the frequently cited model (Pomeroy &
Parkhurst 1977) for the average deterioration rate c of concrete (mm/year):
c = 11.5 k Φ A−1ψ (3)
where ȥ is the coefficient of model uncertainty (optional), k is the factor representing
the proportion of acid reacting (only an estimated value and ranging from 1.0 when
acid formation is slow to 0.3 when it forms rapidly), Ɏ is the flux of H2S (or sulfide
release [g H2S/(m2 hr)]), and A is acid-consumption capability, i.e. the alkalinity of
the concrete, expressed as the proportion of equivalent calcium carbonate (g CaCO3/g
concrete). For granitic aggregate concretes A ranges from 0.17 to 0.24, while for
calcareous aggregates it ranges from 0.9 to 1.1, the equivalent A value for mortar-
lined pipes being 0.4.
The average rate of corrosion represents the total loss of pipe wall over the
total surface area considered. Moreover, the crown of the pipe tends to exhibit a
greater rate of corrosion then the intermediate perimeter. Also, areas of greater
turbulence have higher corrosion. Eq. (3) is thus modified with flux Ɏ analyzed
according to (Tee et al. 2011) and included in the equation used for the determination
of the maximum rate of corrosion cmax (ASCE 2007). The resulting equation reads:

cmax = k ccf ktcf 4,025 ⋅10 −3 k j A −1 [BOD ] (1,07 )


T − 20 §¨ P ·¸ ψ (4)
© (πD − P ) ¹
where kccf is the crown corrosion factor (typically 1.5 to 2), ktcf is the turbulence
corrosion factor (typically 1 to 2 for well-designed drop structures and 5 to 10 for
sharp drops or other turbulent junctions), j is the pH-dependent factor for the
proportion of H2S (e.g. for pH = 7.4 it follows that j = 0.28), P is the wetted perimeter
of the pipe wall (m), D is the diameter of the pipe, [BOD] is the biochemical oxygen
demand concentration (mg/l) and T is the sewage temperature (°C).

© ASCE Rev. 11/2014


CONCREEP 10 222

Software tool
The FReET-D software package, a feasible and user-friendly combination of
analytical models and simulation techniques, has been used for the purposes of the
present paper. Models for carbonation, chloride ingress, reinforcement corrosion,
acid attack and frost attack are available, as well as the model for sulfate attack on
concrete in sewage collection systems shown in previous section. Altogether, 32
models are implemented as pre-defined dynamic-link library functions selected from
the literature. Fully probabilistic safety formats are employed, serving also for the
provision of quantitative information concerning a structure’s safety level. The
uncertainties associated with parameters involved in deterioration processes are
modelled by random variables. Statistical, sensitivity and reliability analyses are
provided. Several features are offered, including parametric studies and Bayesian
updating. Some of the models selected from the literature and originally developed
as deterministic models have been converted into a probabilistic form for the
purposes of this software. For more details see www.freet.cz (Novák et al 2014).

Parametric study
An ad-hoc example of the degradation of concrete sewer pipes induced by
sulfuric acid attack is described here in order to show the potential of the used model
and software. The focus is placed on the influence of two kinds of parameters വ the
wetted perimeter of the pipe wall and the type of aggregate in the concrete. Utilizing
model (4), the stochastic analysis of biogenic sulfate corrosion rate c (mm/year) is
accomplished using the input data set listed in Table 1. The wetted perimeter of the
pipe wall is varied (reflecting the filling of the pipe by 10 % വ 80 %) for two types of
concrete, which is made either from granitic aggregates (according to (ASCE 2007)
the acid-consumption capability value ranges from 0.17 up to 0.24; A = 0.20 is
chosen in present example) or calcareous aggregates (A = 0.9 up to 1.1; A = 1.0 is
used).
The resulting corrosion rates are depicted in Figure 1 (mean ± standard
deviation; the best fitted being the Gumbel max and Lognormal PDFs) demonstrating
the use of the stochastic approach.
The aggregate type appeared to influence concrete degradation to the largest
extent (Belie et al. 2004) – this effect is also clearly evident from the presented
example. Note that in reality the pipe is usually 20 % വ 40 % full, which leads in our
example to a corrosion rate at the crown of about 2 mm/year (granitic aggregate) or
0.4 mm/year (calcareous aggregate). This appears rather realistic with regard to
results mentioned in other sources; e.g. in (ASCE 2007) examples with a rate of 0.7
up to 4 mm/year for calcareous aggregates are given.
An example of reliability and assessment of type (i), Section 3.1 can be
presented together with the service life prediction as follows. Considering a limiting
value for the degraded pipe thickness of 40 mm (i.e. the barrier in Eq. (1), regarding
an SLS case with the design reliability index value ȕ = 1.5), if the reliability analysis
is carried out for granitic aggregate, the corresponding service life is about 18 years.
In the case of calcareous aggregate it would be more than 50 years.

© ASCE Rev. 11/2014


CONCREEP 10 223

Table 1. Input variables.


Mean COV
Variable Unit PDF Reference
value (%)
Uncertainty factor of (JCSS
വ 1 15 Lognormal (2par)
model 2014)
Crown corrosion factor വ Rectangular (ASCE
വ 1.75 വ
kccf (a = 1.5; b = 2.0) 2007)
Turbulence corrosion (ASCE
വ 1.5 വ Deterministic
factor വ ktcf 2007)
Proportion of acid Beta (Bellie et
വ 0.80 5
reacting വ k (a = 0.3; b = 1.0) al. 2004)
TwoBounded (ASCE
i) 0.20 Normal 2007)
Acid-consumption
വ ii) 10 i) a = 0.17; b =
capability വ A
1.00 0.24
ii) a = 0.9; b = 1.1
Factor for the proportion Beta (Pomeroy
വ 0.28 15
of H2S വ j (a = 0; b = 1) 1976)
Biochemical oxygen Typical
mg/l 350 15 Normal
demand - [BOD] value
Typical
Sewage temperature വ T °C 22 10 Normal
value
Diameter of the pipe wall
m 1.8 0.5 Normal
വD
1.158
Wetted perimeter of the (Pomeroy
m വ വ Deterministic
pipe wall വ P 1976)
3.986

MODELING OF STRUCTURAL BEARING CAPACITY

Ring failure
The stochastic structural analysis of buried concrete sewer pipes with the
consideration of biogenic sulfuric acid corrosion is an involved, time dependant task.
Therefore, as a first step, a stochastic study for the strength analysis of the impact of
corrosion on concrete pipes has been performed, based on experimental results
presented in (Karihaloo 1995 വ Chapter 9, with reference to Gustafsson 1985).
The ring failure (or crushing strength fcr) of a non-degraded thin-walled pipe
corresponding to a single force on the top of the pipe (specifically: a linear load, Fig.
2) is studied utilizing concrete fracture theory, namely the FCM (Fictitious Crack
Model), which exhibits a good match to the test summary (2366 samples altogether!).
Note that tests on pipes with a wall thickness (t)/inner diameter (Di) ratio of 1/7 were
considered. Simple linear elasticity proved not to be relevant in this case. In this
paper the crack band model implemented in FEM was used (Atena software ͸

© ASCE Rev. 11/2014


CONCREEP 10 224

ýervenka et al. 2007) to simulate the variation in the crushing strength fcr (Eq. 5,
Karihaloo 1995) of concrete pipes with different wall thickness. The ultimate
experimental top load Pu in Eq. 5 is replaced in this case by the ultimate value from
the nonlinear model.

Figure 1. Rate of corrosion vs. filling of pipe for two types of aggregate.

The obtained numerical results are compared to experiments in Fig. 2. Note


that the 3D Nonlinear Cementitious material model (plain strain idealization;
ýervenka et al. 2007) was employed – tensile strength ft was equal to 4.52 MPa and
characteristic length lch was equal to 380 mm (the same parameters were determined
for experimentally investigated specimens). As lch is defined by Eq. 6, a combination
of fracture energy Gf = 210 N/m and modulus of elasticity E = 37 GPa was used.
These parameters are in close agreement with those of concrete recently developed
and investigated for other applications (ěoutil et al. 2014). For crushing strength the
following holds

6 (Pu / S ) (1 + Di / D0 )
f cr = , (5)
π D0 (1 − Di / D0 )2
where Di (D0) is the inner (outer) diameter of the pipe and Pu/S is the intensity
of the ultimate linear load. Note that t/Di = 1/7. The characteristic length is defined as
EG f
lch = , (6)
ft2
The final model was consequently used to study the influence of the rate of
concrete deterioration on the variation in the crushing strength fcr of concrete pipes
with wall thickness – the results of the study can be seen in Fig. 2. The considered
reference rate of concrete deterioration was 0.5 mm/year at the crown of the pipe (see
scheme in Fig. 2). The study was carried out for 30, 60 and 90 years.

© ASCE Rev. 11/2014


CONCREEP 10 225

Figure 2. (a) Variation in the crushing strength fcr of concrete pipes with wall
thickness – experiments (Karihaloo 1995; black symbols), simulations (white
symbols) and the value of fcr = ft (linear elasticity prediction; horizontal line), (b)
Simulated influence of the rate of concrete deterioration on the variation in the
crushing strength of concrete pipes with wall thickness (from above ͸ 0, 30, 60
and 90 years of degradation).

Moreover, the study was also realised at the stochastic level using the approach
published in (e.g. Novák et al. 2007, Strauss et al. 2008). The rate of concrete
deterioration as well as the following dominant concrete parameters of the material
model were considered to be random variables: Modulus of elasticity, compressive
and tensile strength, and fracture energy. All of the input basic random variables
involved, and the particular set of their statistical parameters (mean value, coefficient
of variation (COV), and probability distribution function (PDF)) are summarized in
Tab. 2. Statistical correlation among the input basic random variables was also
considered (employing simulated annealing – Novák et al. 2007); see Tab. 3.
Correlation coefficients were set up based on previous experiments (e.g. Novák et al.
2007, ěoutil et al. 2014). The Latin Hypercube Sampling method (Novák et al. 2014)
was used and 16 simulations of crushing strength fcr were performed for selected
concrete pipe profiles and degradation times (0, 30, 60 and 90 years). The obtained
results – values of crushing strength fcr and their statistical parameters – are shown in
Fig. 3 and Tab. 3, and can be used as inputs in the subsequent reliability and
durability analysis of concrete pipes.
The results summarized in Tab. 3 show an interesting trend – the value of fcr for the
pipe with DN 400 (wall thickness 60 mm) after 30 years of degradation is nearly the
same as for the non-degraded pipe DN 600, etc. – see the bold numbers in the
diagonals.
Table 2. Basic random variables of concrete deterioration and dominant
material parameters.
Variable [Unit] Mean value COV PDF
Modulus of elasticity E [GPa] 37 0.1 Log-normal
Compressive strength fc [MPa] 69.5 0.1 Log-normal
Tensile strength ft [MPa] 4.52 0.12 Log-normal
Fracture energy Gf [J/m2] 210 0.2 Normal
Deterioration rate of concrete [mm/years] 0.5 0.33 Log-normal

© ASCE Rev. 11/2014


CONCREEP 10 226

Table 3. Correlation coefficients


between dominant material
parameters.
Variable E fc ft Gf
E 1 0.9 0.7 0.37
fc 1 0.9 0.6
ft 1 0.9
Gf Sy 1
m.

Figure 3. Crushing strengths fcr for selected wall thicknesses (60, 100 and 140
mm) and deterioration times – 0 (a), 30 (b), 60 (c) and 90 (d) years - obtained
using a stochastic model.

Table 4. Basic statistical parameters of crushing strengths for selected wall


thickness and times of deterioration.
t = 0 years t = 30 years t = 60 years t = 90 years
mean [MPa] 8.91 7.26 6.89 4.99
COV 12.8 15.2 13.6 39.0
DN400
PDF Weibull max Gumbel Gumbel Weibull
(3 par) Max. EV I Max. EV I min (3 par)
Mean [MPa] 7.55 6.56 6.06 5.71
DN700 COV 12.2 11.5 13.8 10.7
PDF Weibull max Weibull max
(3 par)
Normal Rayleigh (3 par)
Mean [MPa] 6.90 6.42 5.78 5.25
DN1000 COV 12.1 12.2 12.4 16.3
PDF Weibull min Gumbel Gumbel
Normal (3 par) Max. EV I Max. EV I

Buried pipe analysis


The strength analysis of buried concrete sewer pipe behaviour with
consideration given to material nonlinearity, contact with different soil layers, the
MIC degradation of the concrete, and the involved uncertainties, as well as the
assessment of reliability and the prediction of service life, are all complex tasks. They
can be solved employing materially nonlinear FE analysis combined with a suitable
numerical simulation technique in a similar manner to that described in the previous
section. Note that different limit states and relevant reliability index values can be
assessed in this way.

Creep effect
The behaviour of the buried sewer pipe is time-dependant due to the MIC
degradation effect. Evidently the creep of concrete is an additional time-variant effect
which can play a considerable role. It can be incorporated into the above described
approach in a simplified way: at each time step a creep strain value (associated with
actual compressive stress) is added to the strain value due to mechanical loading in

© ASCE Rev. 11/2014


CONCREEP 10 227

each element. The statistical parameters of the creep strain can be gained using a
special FreET module which employs e.g. the B3 model.
Note that the approach described briefly in the last two sub-sections is part of
work which is currently ongoing; some results will be reported at the conference.

CONCLUSION

The microbiologically induced corrosion of reinforced concrete sewers and


manholes is a problem with significant global repair costs. When corrosion is
sufficiently advanced it can lead to diminished service life or to structural failures.
Therefore, for the life cycle management of sewage facilities proper and effective
tools are needed; they may serve in this respect in several ways, e.g.:
(i) Utilizing suitable models, the assessment of concrete deterioration rate affected
by different material or operational parameters can be studied, e.g. in the form of
parametric studies, service life prognosis, etc.
(ii) Using a model and considering a given time period, material and operational
parameters, the thickness of a buried concrete pipe weakened due to corrosion is
specified. Based on such a degraded structural configuration, the reliability of the
bearing capacity of the pipe can be assessed (durability, SLS or ULS).
(iii) In a reciprocal manner with regard to variant (ii), first the bearing capacity of the
buried pipe is determined, i.e. the necessary (minimal) cross sectional and
material characteristics. Then, utilizing a relevant degradation model, a service
life prognosis can be performed.
This can help managers make more realistic decisions concerning the safe and
cost-effective operation of pipes during their service life.

Acknowledgement
This work has been supported by projects No. 14-10930S and 13-22899P
awarded by the Czech Science Foundation and by project CZ.1.07/2.3.00/30.0005 –
“Support for the creation of excellent interdisciplinary research teams at Brno
University of Technology”.

REFERENCES
ASCE (American Society of Civil Engineers). (2007) Manuals and Reports of
Engineering Practice, No. 60, 2007, Gravity Sewers, NY, USA.
ýervenka, V., Jendele, L. and ýervenka, J. (2007). “ATENA program documentation
– Part 1: theory.” Prague, Czech Republic: Cervenka Consulting.
De Belie De, N. et al. (2004). “Experimental research and prediction of the effect of
chemical and biogenic sulphuric acid on different types of commercially
produced concrete sewer pipes.” Cem. Concr. Res., 34, 2223–2236.
Fernardes, I. et al. (2012). “Identification of acid attack on concrete of a sewage
system.” Mat. Struct., 45(3), 337–350.
fib Model Code 2010 (2012). fib Bulletins No. 65 and 66. International Federation for
Structural Concrete, Lausanne, Switzerland.
ISO 13823 (2008). General Principles on the Design of Structures for Durability.

© ASCE Rev. 11/2014


CONCREEP 10 228

ISO 16204 (2012). Durability – Service life design of concrete structures.


JCCS (Joint Committee on Structural Safety) (2014). Accessible from
<http://www.jcss.byg.dtu.dk/Publications/Probabilistic_Model_Code.aspx >.
Karihaloo, B. L. (1995). Fracture Mechanics and Structural Concrete. Longman.
Neville, A. (2004). “The confused world of sulfate attack on concrete.” Cem. Concr.
Res., 24, 1275–1296.
Novák, D., VoĜechovský, M., Lehký, D., Bergmeister, K., Pukl, R., and ýervenka V.
(2007). “Stochastic nonlinear analysis of concrete structures – Part I.” Proc. of
ICASP10, Tokyo, ISBN 978-0-415-45211-3.
Novák, D., VoĜechovský, M. and Teplý, B. (2014). “FReET – Software for the
statistical and reliability analysis of engineering problems and FReET-D:
Degradation Module.” Advances in Engineering Software, 72, 179–192.
Pomeroy, R. D. and Boon, A. G. (1976). The problem of hydrogen sulphide in sewers.
Clay Pipe Development Association, Ltd.
Pomeroy, R. D. and Parkhurst, J. D. (1977). “The forecasting of sulfide build-up rates
in sewers.” Progr. Water Technol., 9(3), 621–628.
RILEM Report rep043 (2013). Publications on Durability of Reinforced Concrete
Structures under Combined Mechanical Loads and Environmental Actions: An
Annotated Bibliography. Edited by Yao Yan, Wang Ling, Wittmann Folker H.
ěoutil, L., Lehký, D., Šimonová, H., Kucharczyková, B., Keršner, Z., Novák, T.,
Zimmermann, T., Strauss, A. and Krug B. (2014). “Experimental-computational
determination of mechanical fracture parameters of concrete for probabilistic life-
cycle assessment.” Proc. of IALCCE 2014, Tokyo, 801വ807.
Sakar, S., Mahadevan, S., Meeussen, J.C.L., Sloot, H. van der and Kosson, D.S.
(2010). “Numerical simulation of cementitious materials degradation under
external sulfate attack.” Cem. Concr. Com., 32, 241–252.
Stein, D. (1999). Instandhaltung von Kanalisationen. 3rd edition, Berlin: Ernst, p.
141. ISBN 3-433-01315-2. (in German)
Strauss A, Bergmeister K, Hoffmann S, Pukl R and Novak D. (2008). “Advanced
life-cycle analysis of existing concrete bridges.” J. Mater. Civil Eng. (ASCE)
20(1), 9–19.
Sun, Ch., Chen, J., Zhu, J., Zhang, M. and Ye, J.(2013). “A new diffusion model of
sulfate ions in concrete.” Constr. Build. Mater., 39, 39–45.
Tee, K. F., Li, CH. Q. and Mahmoodian, M. (2011). “Prediction of time-variant
probability of failure for concrete sewer pipes.” Proc. of XII DBMC, Porto, Vol. I,
447–454.
Teplý, B. and VoĜechovská, D. (2012). “Reinforcement corrosion: Limit states,
reliability and modelling.” J. of Advanced Concrete Technology, 10, 353–362.
Wells, T. and Melchers, R. E. (2014). “An observation-based model for corrosion of
concrete sewers under aggressive conditions.” Cem. Concr. Res., 61-62. 1–10.
Yamamoto, D., Hamada, H., Sagawa, Y and Md Noor, N. (2013). “Diagnostic of load
bearing capacity of sewage pipes deteriorated by sulphate attack.” Proc. of
CONSEC13, 1280–1288.
Yuan, H., Dangla, P., Chatellieur, P. and Chaussadent, T. (2013). “Degradation
modelling of concrete submitted to sulphuric acid attack.” Cem. Concr. Res., 53,
267–277.

© ASCE Rev. 11/2014


CONCREEP 10 229

Coupled Effects of Static Creep, Cyclic Creep and Damage on WKHLong-Term


Performance of Prestressed Concrete Bridges: A Case Study Based on
Rate-Type Formulation

Qiang Yu1 and Teng Tong2

1
Department of Civil and Environmental Engineering, University of Pittsburgh, PA
15261. E-mail: qiy15@pitt.edu
2
Department of Civil and Environmental Engineering, University of Pittsburgh, PA
15261. E-mail: tet16@pitt.edu

Abstract
One of the challenges in bridge engineering is to predict the long-term deflection and
damage accumulation in large-span prestressed concrete girders during their lifespan.
For prestressed concrete bridges, a salient by-product of time-dependent deformation
resulting primarily from concrete creep is unexpected stress redistribution, which
triggers damage and cracking in the critical structural components. To deepen the
understanding of the coupled effects of deflection and damage on bridge
performance, a case study is carried out in this investigation, in which a large-span
prestressed concrete bridge is modeled in ABAQUS based on its blueprints available.
To capture the concrete behavior during service, a constitutive law formulated based
on a damage model is employed to approximate the nonlinear softening after concrete
cracking. This formulation is further integrated with a rheological model resting on
Kelvin units for concrete creep. Using a rate-type algorithm, the long-term behavior
of the bridge is probed.
INTRODUCTION
With the increasing demands in construction sustainability, a longer lifespan, i.e.,
over 100 years, is now generally expected for critical prestressed concrete girders. A
serious challenge to this goal is the low predictive capacity of the current design
models for bridge long-term performance, a complex outcome of time-dependent
interactions between concrete creep, shrinkage, steel relaxation, and deterioration
processes (Strauss et al. 2013, Wendner et al. 2010), coupled with other physical and
mechanical influences.

© ASCE Rev. 11/2014


CONCREEP 10 230

In a recent survey (Bažant et al. 2012), deflections exceeding design


expectation are found prevalent in large-span prestressed concrete girders. The
excessive deflection triggers unexpected cracking in concrete members, and thus
significantly compromises the safety and serviceability of prestressed concrete
bridges. A primary source of the inaccuracy is identified to be the inadequacy existing
in the current concrete creep models used in bridge design (Bažant et al. 2013,
Wendner et al. 2015). In addition, lack of coupling between deflection and other
important physical and chemical processes further exacerbates the prediction
accuracy. It is found the popular, yet primitive formulation based on integral form
used in the current creep structural analyses, which is incapable of taking into account
time-independent phenomena (e.g., concrete cracking), leads to substantial
underestimate of long-term deflection (Yu et al. 2012).
To overcome these obstacles in creep structural analysis and bridge design,
rate-type formulation residing on improved creep models was recently developed and
implemented in long-term deflection analysis (Bažant et al. 2013, Wendner et al.
2015, Yu et al. 2012). In the rate-type formulation, concrete creep is approximated by
a rheological model and continuous spectrum method is employed to uniquely
identify the viscoelastic parameters based on the given creep compliance function.
A prominent advantage of rate-type formulation is that it allows the creep
structural analysis to be coupled with other time-independent phenomena, especially
concrete cracking trigged by stress redistribution. Unfortunately, the research effort to
explore the coupled effects of time-dependent deformation and concrete damage is
limited in bridge structural analysis, especially for long-term performance prediction.
The objective of this study is to build a unified rate-type formulation to capture the
coupled effects of concrete creep and cracking on bridge long-term performance. A
case study based on a large-span bridge will be carried out and the results obtained
from FEM simulation will be compared with the in-situ measurements.
GENERAL FRAMEWORK
Under service load, concrete creep can be deemed as an ageing linear viscoelastic
phenomenon, which can be accurately approximated by a rheological model (Bažant
and Prasannan 1989 a, b, Yu et al. 2012). Although Maxwell units are equivalently
effective, Kelvin units are popularly used in creep structural analysis due to their
direct correlation with the parameters observed in creep tests. As shown in Fig. 1,
Kelvin units characterized by moduli Di and retardation time τi are coupled in a series
connection.
Unexpected deformation causes stress redistribution, which triggers damage
and cracking in concrete. To capture the nonlinear softening of concrete after
cracking, the rheological model is further enriched by constitutive laws hinging on
isotropic damage. When stress exceeds the threshold (e.g., concrete tensile strength),
damage happens and it influences the behaviour of Kelvin units. Correspondingly, a
damage factor will be assigned to Di in an isotropic manner.

© ASCE Rev. 11/2014


CONCREEP 10 231

In addition to static creep which is characterized by strain growth under


sustained stress, cyclic creep due to cyclic loading contributes to the time-dependent
deformation. For bridges under moderate traffic loads or bridges of a short lifespan
(e.g., less than 30 years), cyclic creep does not play an important role (Bažant and
Hubler, 2014). However, for large-span critical bridges of an extended lifespan (over
100 years), the contribution of cyclic creep cannot be ignored.
As shown in Fig. 1, the total strain of concrete εij can be conceptually
expressed as
ε ij = (ε ije ) + (ε ij′′ + ε ijcc + ε ijsh )
N 

ε iji ε ijt

where ε iji is the instantaneous strain after loading, and ε ijt is its time-dependent
counterpart consisting of inelastic strain induced by concrete static creep ε ''ij , cyclic
creep ε ijcc and shrinkage ε ijsh . In this study, autogeneous shrinkage is neglected and
only the drying shrinkage resulting from moisture loss is taken into account.

Figure 1. One dimensional illustration of the proposed unified concrete model

RATE-TYPE FORMULATION
Within the service stress level, concrete static creep is assumed to follow the
constitutive law of ageing linear viscoelasticity (Jirásek and Bažant 2002, Bažant et
al. 2012 a, b). Due to concrete ageing, the compliance function J(t, t’), serving as
kernel in Volterra integral form, is not of a convolution type. Therefore, it may be
approximated by a rheological model in the rate-type formulation so as to capture the
strain growth under a general stress history σ (t).
If Kelvin units are employed, the compliance function can be approximated as
(Hardy and Riesz 2013, Schapery 1962, Jirásek and Bažant 2002):
1 M
1 t − t′
J (t , t ′) ≈ +¦ [1 − exp(− )]
E (t ) i =1 Di (t ′) τi
where t is current time and t’ is the loading age. To avoid thermodynamically
inadmissible Di obtained through data fitting, continuous spectrum method is adopted

© ASCE Rev. 11/2014


CONCREEP 10 232

to uniquely and efficiently identify the compliance spectrum Ai. Utilizing Laplace
transformation inversion supplemented by Widder’s approximate inversion formula
(Widder 1971), the analytical solution of Ai for a given compliance function can be
attained; see Yu et al. 2012.
Damage in concrete can be approximated by a number of forms based on
different formulation. Here a damage model resting on an isotropic damage variable
Φ will be employed (Kachonov, 1958). Using the strain equivalence hypothesis and
generalized Hook’s law, the effective (undamaged) stress tensor σij can be expressed
as:
σ ij = Eijkl (t )ε kle = Eijkl (t )ε kle
where E ijkl (t) is the fourth-order effective (undamaged) isotropic elasticity tensor,
expressed as a function of concrete age t. After concrete cracking, concrete stress-
strain relation can be described by a simplified softening law as:
σ ij = (1 − Φ)σ ij = (1 − Φ) Eijkl (t )ε ije
here the evolution of damage variable Φ follows an isotropic damage rule hinging on
the critical principle strain of concrete. In the formulation, this damage variable will
be assigned to all Kelvin units based on the assumption that same degree of damage
happens in each Kelvin unit.
For the cyclic creep, a power law-type equation similar to Paris law will be
used. The development of cyclic creep after N cycles is approximated as (Bažant and
Hubler, 2014):
Δσ m
ε cc = C1σ ( ) N
fc′
Here f c' is the compressive strength of concrete, Δσ is the amplitude of stress
variation and σ is the average stress in concrete. Note that ε cc depends on both σ
and N linearly, which agrees with the test measurements available in literature and is
convenient for structural analysis. In this study, the exponent value m is set as 4, and
C1 is about 46×10-6.
For a rate-type formulation, the 3D quasi-elastic stress-strain incremental
relation at any time step can be written as:
Δσ = E′′Δε e
where E '' is the effective incremental modulus considering viscoelasticity and
concrete cracking. The increment of the effective elastic strain Δεe is:
Δε ije = Δε ij − Δε ijcc − Δε ijsh − Δε ij′′

© ASCE Rev. 11/2014


CONCREEP 10 233

It can be seen that in the rate-type formulation, the creep calculation is


coupled with other time-dependent processes in the structural analysis. The general
algorithm to be implemented in FEM can be illustrated in Fig.2.

Figure 2. Overall algorithm for rate-type formulation.


CASE STUDY BASED ON A LARGE-SPAN BRIDGE
To explore the coupled effects of time-dependent deformation and concrete cracking,
a case study based on a large-span prestressed concrete bridge is carried out. Utilizing
the information available in the blueprints, a 3D model of the bridge is built in
ABAQUS; see Fig. 3. As a segmentally erected bridge, the bridge consists of 138
cast-in-situ segments and its main span is over 270 m. In order to focus on the post-
construction behavior, the camber generated in segmental construction, which is
varied for different creep models, will not be considered in the deflection comparison.

© ASCE Rev. 11/2014


CONCREEP 10 234

Figure 3. 3D finite element model of a large-span bridge (half-model)


In the modelling, ACI (ACI, 2008), fib MC2010 (fib, 2012) and B4 (Wendner et al.
2015) models are employed for concrete creep and shrinkage. Among these models,
ACI model is purely empirical and the only intrinsic parameter employed to represent
concrete characteristics in its compliance function is concrete strength f c' . The fib
MC2010 is similar to its old version, namely CEB-FIP MC90 model, except for
introducing a split into basic and drying creep with a logarithmic time function to
capture the long-term behaviour of creep observed in long-term creep tests (Brooks
1984, Brooks 2004). B4 model, based on the solidification theory (Bažant and
Prasannan 1989 a, b, Bažant and Baweja 2000), split concrete creep into basic creep
and drying creep, the former being unbounded and consisting of short-term strain,
viscous strain and a flow term, while the latter being bounded and related to moisture
loss. Besides f c' , multiple intrinsic parameters are employed in B4 model to represent
concrete composition characterized by water-cement ratio (w/c), aggregate-cement
ratio (a/c), and cement content (c).

(a) Linear time scale (b) Logarithmic time scale


Figure 4. Comparison of deflection at midspan 7 years after construction.
As for prestressing steel, a bilinear law featured with Young’s modulus E and
yield strength fy is used in the simulations. Its relaxation is modelled based on the
CEB relaxation formulas (CEB 1990). To simplify the calculation of cyclic creep, a
normalized cyclic traffic load based on annual traffic report is used in the modelling.

© ASCE Rev. 11/2014


CONCREEP 10 235

(a) 3 years (b) 4 years

(c) 5 years (d) 6 years


Figure 5. Comparison of the deflection profile.
Fig 4 shows the comparison between the in-situ measurements and predictions
from ACI, fib MC 2010 and B4 models without considering damage and cyclic creep.
After 7 years, the predicted deflection at midspan is about 100, 110 and 140 mm for
the ACI, fib MC 2010 and B4 models, respectively. While, the measured deflection is
about 210 mm. The substantial deviation indicates that the neglect of damage and
cyclic creep of the concrete will lead to significant underestimate of long-term
deflection, especially for bridges carrying heavy traffic load.
If concrete damage and cyclic creep are considered, remarkable improvement
can be achieved in deflection prediction. In Fig. 5, the simulated deflection profiles of
the bridge are plotted. From Figs 5(a) to (d), comparison is made with the in-situ
measurements obtained after 3-, 4-, 5- and 6-year service respectively. Clearly, the
prediction based on all models agrees well with the measured deflection history.
CONCLUSIONS
In this study, a unified concrete model coupling the time-dependent behaviour of
concrete (e.g. quasi-static and cyclic creep, shrinkage, etc.) with concrete damage is
developed and its effectiveness is demonstrated in a case study. Based on the FEM
simulation, the following conclusions can be drawn:

© ASCE Rev. 11/2014


CONCREEP 10 236

(1) For the prestressed concrete bridges carrying heavy traffic flow cyclic creep of
concrete must be taken into account in bridge design. The neglect of cyclic creep
may lead to unrealistic long-term deflection prediction.
(2) Due to the time-dependent deformation, cracking will happen in concrete as a
consequence of stress redistribution. The interaction between concrete damage
and creep will augment the time-dependent deflection of prestressed bridges.
(3) To capture the coupled effect of damage and time-dependent deformation, rate-
type formulation is needed. In addition to cyclic creep and concrete cracking, it
can be further extended to incorporate other physical and chemical processes,
e.g., corrosion.

REFERENCES
ACI Committee (2008). Building Code Requirements for Structural Concrete (ACI
318-08) and Commentary. Farmington Hills: American Concrete Institute.
Bažant, Z. P., and Prasannan, S. (1989 a). “Solidification theory for concrete creep II:
verification and application.” Journal of Engineering mechanics, 115(8),
1704-1725.
Bažant, Z. P., and Prasannan, S. (1989 b). “Solidification theory for concrete creep I:
formulation.” Journal of engineering mechanics, 115(8), 1691-1703.
Bažant, Z. P., and Baweja, S. (2000). “Creep and shrinkage prediction model for
analysis and design of concrete structures: Model B3.” ACI SPECIAL
PUBLICATIONS, 194, 1-84.
Bažant, Z. P., Hubler, M. H., and Jirásek, M. (2012). “Improved estimation of long-
term relaxation function from compliance function of aging concrete.” ASCE
J. of Engrg. Mech, 139 (2), 146–152.
Bažant, Z. P., Yu, Q., & Li, G. H. (2012 a). “Excessive long-time deflections of
prestressed box girders. I: Record-span Bridge in Palau and other paradigms.”
Journal of Structural Engineering, 138(6), 676-686.
Bažant, Z. K. P., Yu, Q., & Li, G. H. (2012 b). “Excessive long-time deflections of
prestressed box girders. II: Numerical analysis and lessons learned.” Journal
of Structural Engineering, 138(6), 687-696.
Bažant, Z. P., Hubler, M. H., and Jirásek, M. (2013). “Improved Estimation of Long-
Term Relaxation Function from Compliance Function of Aging
Concrete.” Journal of Engineering Mechanics, 139(2), 146-152.
Bažant, Z. P., & Hubler, M. H. (2014). “Theory of cyclic creep of concrete based on
Paris law for fatigue growth of subcritical microcracks.” Journal of the
Mechanics and Physics of Solids, 63, 187-200.
Brooks, J. J. (1984). “Accuracy of estimating long-term strains in concrete.”
Magazine of Concrete Research, 36(128), 131-145.
Brooks, J. J. (2005). “30-year creep and shrinkage of concrete.” Magazine of concrete
research, 57(9), 545-556.

© ASCE Rev. 11/2014


CONCREEP 10 237

Fédération Internationale du Béton (2012). International Federation for Structural


Concrete Special Activity Group New Model Code. Lausanne: International
Federation for Structural Concrete.
Hardy, G. H., and Riesz, M. (2013). The general theory of Dirichlet's series. Courier
Corporation.
Jirásek, M., and Bažant, Z. P. (2002). Inelastic analysis of structures. NewYork: John
Wiley & Sons.
Kachonov, L.M. (1958). “On the creep fracture time.” Izvestiya Akademii Nauk USSR
Otd. Tech. 8, 26–31 (in Russian).
Schapery, R. A. (1962). “Approximate methods of transform inversion for
viscoelastic stress analysis.” In Proceedings Fourth US National Congress of
Applied Mechanics, 2, 1075-1085.
Strauss, A., Wendner, R., Bergmeister, K., and Costa, C. (2013). “Numerically and
Experimentally Based Reliability Assessment of a Concrete Bridge Subjected
to Chloride Induced Deterioration.” Journal of Infrastructure Systems, 19(2),
166-175.
Wendner, R., Strauss, A., Guggenberger, T., Bergmeister, K., and Teply, B. (2010).
“Approach for the Assessment of Concrete Structures Subjected to Chloride
Induced Deterioration.” BETON- STAHLBETONBAU, 105(12), 778-786.
Wendner, R., Hubler, M.H., and Bažant, Z.P. (2015). “Statistical justification of
model B4 for multi-decade concrete creep using laboratory and bridge
databases and comparisons to other models.” Materials and Structures, 1-19.
Widder, D. V. (1971). An introduction to transform theory (Vol. 42). Academic Press.
Yu, Q., Bažant, Z. P., and Wendner, R. (2012). “Improved Algorithm for Efficient
and Realistic Creep Analysis of Large Creep-Sensitive Concrete
Structures.” ACI Structural Journal, 109(5), 665-675.

© ASCE Rev. 11/2014


CONCREEP 10 238

Creep of Concrete and Its Instant Nonlinear Deformation in WKH


&DOFXODWLRQof Structures

Rudolf Sanjarovskiy1; Tatyana Ter-Emmanuilyan2; and Maksim Manchenko3


1
Department of Theoretical Mechanics, Saint-Petersburg State University of
Architecture and Civil Engineering (SPSUACE), 2-nd Krasnoarmeiskaya St., 4, St.
Peterburg 190005, Russia. E-mail: milasanj@gmail.com
2
Department of Petroleum Engineering, Kazakh-British Technical University
(KBTU), Tolebi St., 59, Almaty 050000, Republic of Kazakhstan. E-mail:
tanya_ter@mail.ru
3
FSUE "Krylov State Research Center", Moscow Highway, 44, St. Petersburg
196158, Russia. E-mail: salsa87@bk.ru

Abstract

This article is a continuation of the series of publications of authors to


produce instant nonlinear properties of the concrete. The theory, which is a
substantial development of the methods for calculating the long resistance of
reinforced concrete structures, allows almost exactly the instant nonlinear properties
of the concrete.

1. INTRODUCTION

Concrete is essentially nonlinear structural material. The diagram


instantaneous compression σ − ε M (Figure 1) has a decreasing portion bounded by
the limiting deformation ε B 2 . The parameters of this nonlinear diagram normalized to
Eurocode 2 in the section on the calculation of reinforced concrete structures.
In the theory of creep of concrete (linear, nonlinear), instead of a real
nonlinear diagrams used fictitious linear diagram that satisfies Hooke's law (Figure
1). This diagram makes calculations in the two types of errors, corresponding to the
four points of the fictitious 1-4, accompanied by a realistic point of M. For example,
at a given instant deformation ε M fictitious stress σ f = ε M EB is significantly greater
than the actual stress σ for a given real stress ı fictitious elastic deformation
ε e = σ / EB is much less real deformation ε M . Replacing the actual nonlinear
diagram σ − ε M makes large errors in the calculation of total deformations during
prolonged uploading designs.
This approach is justified misleading statement that "experiments
instantaneous deformation of concrete, even at high levels of loading are linearly
dependent on the stress".

© ASCE
CONCREEP 10 239

Figure 1 - The diagram σ − ε M instantaneous deformation of concrete

In the theory of creep of concrete (linear, nonlinear), instead of a real


nonlinear diagrams used fictitious linear diagram that satisfies Hooke's law (Figure
1). This diagram makes calculations in the two types of errors, corresponding to the
four points of the fictitious 1-4, accompanied by a realistic point of M. For example,
at a given instant deformation ε M fictitious stress σ f = ε M EB is significantly greater
than the actual stress σ ; for a given real stress ı fictitious elastic deformation
ε e = σ / EB is much less real deformation ε M . Replacing the actual nonlinear
diagram σ − ε M makes large errors in the calculation of total deformations during
prolonged uploading designs.
This approach is justified misleading statement that "experiments
instantaneous deformation of concrete, even at high levels of loading are linearly
dependent on the stress".
As is known, the curved diagrams concrete have obtained experimentally for
a long time ago. For example, Tal K. in 1955 revealed the real diagram σ − ε M in
concrete by testing prisms reinforced with high-strength steel wire. As can be seen
from Fig. 1, the total instantaneous deformation in the experiments is the sum of the
linear and nonlinear component: ε M = ε l + ε n . With prolonged uploading total
deformation consists of two components: the total instantaneous and creep:
ε = ε ɦ + ε c . The total instantaneous deformation ε M is determined for the time
measured in minutes (at the Alexandrovsky S. listed 4 min.). Creep deformation is

© ASCE
CONCREEP 10 240

manifested in the time measured for days and years, which creates problems for their
joint consideration.

2. BASES OF THE METHOD

In the traditional theory of hereditary creep deformation nonlinear component


İɧ is not considered, so the total deformation is composed of elastic deformation İɥ
and creep deformation: ε = ε l + ε c .These theories include the creep equation: Kelvin,
Boltzmann (1887), Volterra (1913), Maslow-Harutyunyan (1952). In this regard,
Gvozdev in 1955, on the basis of experiments Borishansky M. pointed out that the
traditional theory of creep is not suitable for the theory of reinforced concrete. This
theory does not reflect the experimentally observed rapid leakage of creep
deformation at the time of observation, close to the time of loading of the samples.
The initial portions of the creep curves constructed from these theories creep, do not
have the characteristic outlines of curves at time IJ, close to the time t.
The error lies in the fact that the nonlinear part of the instantaneous
deformation is transferred to her inappropriate discharge creep deformation and
formally joins them. This approach requires compliance with the relevant
mathematical transformations that have not been fulfilled. The above translation is
required to the traditional form of the equation of linear viscoelasticity

σ (t ) t
∂ 1
t

ε (t ) = − ³ σ (τ ) dτ − ³ σ (τ ) C ( t ,τ ) dτ , (1)
E (t ) τ1 ∂τ E (τ ) τ1 ∂τ

with instantaneous elastic deformation.


Retaining the first two terms describing the elastic deformation, the model
recommends transfer to account for deformation ε n (Figure 1) to clarify the kind of
measures the creep C(t, IJ). Since deformation ε n increases nonlinearly with
increasing stress ı, and then measure the nonlinear creep should depend on the stress
Cn(ı, t, IJ). Unfortunately, this requirement was not taken into account by researchers.
Accounting instantaneous deformation creep of concrete was performed for
the first time Yashin A. and Katin N. in 1959, then Alexandrovsky S. and many
foreign scientists. We found that in the entry (1) contains a number of inaccuracies,
introducing an error in the calculation results.
Consider first elastic deformation in equation (1). Generally linear (potential)
forces allows us to find the speed of the elastic deformation

1 d 1
εe (τ ) = σ (τ ) + σ (τ )
E (τ ) dτ E (τ )

and its significance

© ASCE
CONCREEP 10 241

σ (τ 1 ) t 1
t
d 1
ε e (τ ) = + ³ σ (τ ) dτ + ³ σ (τ ) dτ .
E (τ 1 ) τ1
E (τ ) τ 1
dτ E (τ )

After transformations we have ε e ( t ) = σ ( t ) E ( t ) . It follows that the second


term in (1) is superfluous, and used form of the principle of superposition is
incorrect:

σ (τ 1 ) t 1
ε e (τ ) =
E (τ 1 ) τ³ E (τ )
+ dσ .
1

The error is a loss of the deformation portion, which corresponds to the rate
of change of stiffness. The same error exists in the last integral term. Under the
conditions of nonlinear creep further this loss leads to the formulation of the
principle of superposition strange that violates not only the principles of Newtonian
mechanics, but also the conditions of the affine scaling experimental creep curves.
Consider the third term in (1), and write with the help of that part of the linear creep,
which is called instantaneous.
t

t
∂C (σ , t ,τ ) t
∂C (σ , t ,τ )
ε n ( t ) = ³ σ (τ ) Cn (σ , t ,τ ) dτ = ³ σ (τ ) σ (τ ) n dτ + ³ σ (τ ) n dτ .
τ1 ∂τ τ1 ∂σ τ1 ∂τ

However, the calculation of the instantaneous creep does not operate, and the
integral is usually written in the form
t

ε n ( t ) = ³ σ (τ ) Cn ( t ,τ ) dτ . (2)
τ1 ∂τ

Here are the values ε n (t ) , linearly depends on the stresses, that corresponds
neither to experiment to find ε n (t ) , or the data in Fig. 1 of the nonlinear coupling
ε n and ı.
To describe Cn(t,IJ) used a variety of complex formulas that do not meet the
obvious experimental data. For example, in the approximation of the diagram σ − ε M
square parabola, we have the exact value of the deformation ε n = β 2σ 2 :

t

ε n ( t ) = β 2σ 2 ( t ) = ³ σ (τ ) Cn ( t ,τ ) dτ ,
τ1 ∂τ

which should correspond to the right side, including its function C(t, IJ), that it is
impossible to perform traditional records, for example, at Alexandrovsky S.

© ASCE
CONCREEP 10 242

1 − A2 e −γτ −γ ( t −τ )
C ( t ,τ ) = ψ (τ ) −ψ ( t ) e + Δ (τ ) ª¬1 − e −γ ( t −τ ) º¼ .
1 − A2 e −γ t

Complex formulas in the description of C(t, IJ), designed to take into account
the instantaneous creep significantly increase the order of the corresponding
differential equation of creep of concrete. This complicates the solution of practical
problems of ordinary calculation of reinforced concrete structures.
Update the last term in (1), using the properties of potential forces under
creep conditions (Sanzharovsky R., 2014). Find the velocity of creep deformation-

∂C ( t ,τ ) ∂C ( t ,τ )
εc ( t ,τ ) = σ (τ ) C ( t ,τ ) + σ (τ ) + σ (τ ) ,
∂τ ∂t

its magnitude and

t t
∂C ( t ,τ ) t
∂C ( t ,τ )
ε c ( t ) = σ (τ 1 ) C ( t ,τ 1 ) + ³ σ (τ ) C ( t ,τ ) dτ + ³ σ (τ ) dτ + ³ σ (τ ) dτ .
τ1 τ1 ∂τ τ1 ∂t

Finally, we have after transformation

t
∂C ( t ,τ )
ε c ( t ) = ³ σ (τ ) dτ . (3)
τ1 ∂t

The last term in (1) and the value of (3) in linear creep may not differ from
each other only by using the difference of the nuclei. This fact characterizes the
corresponding superposition principle, as well as the correctness of its further
application in the nonlinear creep.
These unauthorized principles and errors have a significant impact on the
results of the calculations of reinforced concrete structures creep. They also show
that in the conventional form of (1) the equation for creep regulation and mass use in
reinforced concrete structures is unacceptable.
Previously, the authors of this paper (Sanzharovsky R., 2013) proposed a
method of accounting instantaneous nonlinearity of concrete creep. According to the
hypothesis of linear concrete creep basic equation by creep can be written as
t

ε ( t ) = f 2 ª¬σ ( t ) º¼ + ³ f1 ª¬ε M (τ ) º¼ C ( t ,τ ) dτ , (4)
τ1 ∂t

which f1 ª¬ε M (τ ) º¼ , f 2 ª¬σ ( t ) º¼ represent the direct and inverse function of


nonlinear diagrams instantly σ − ε M concrete parameters which change over time.
For example, the use of concrete for nonlinear diagram with the descending
parts in the form proposed by Eurocode 2

© ASCE
CONCREEP 10 243

aε M + δε M2 a σ + δ 0σ 2 1
σ = f1 ( ε M ) = ; ε M = f 2 (σ ) = 0 ; a = EB ( t ) ; a0 = ;
c1 + gε M c0 + g0σ EB ( t )

Rδ ( t ) EB ( t ) ε B1 − 2 RB ( t )
δ =− ; g= ; c1 = c0 = 1 and so on.
ε 2
B1 RB ( t ) ε B1

In its simplest form using square parabola

σ = f1 ª¬ε M ( t ) º¼ = EB ( t ) ε M − A2 ( t ) ε M2 , (5)
1
εM = σ + B2 ( t ) σ 2 = f 2 ª¬σ ( t ) º¼ ,
EB ( t )

where Eɜ(t), A2(t), B2(t) - known functions.


To measure the creep of concrete C (t, IJ) can use any sentence of the famous
scientists.
For the calculation of the valuation of the most common structures
(reinforced concrete beams and columns), the average consumer friendly (
Sanzharovsky R., 1978), the function C(t, IJ) should be presented in a degenerate
form. In this case, equation (4) reduces to a differential form, and final
recommendations for designers are represented in a simple tabular form. When a
degenerate form of the type C ( t ,τ ) = θ (τ ) ª¬1 − e −γ (t −τ ) º¼ proposed by Prokopovich I.
and Karapetyan K. the differential equation of creep of concrete will be first order.
Here we use a measure of creep in the form proposed McHenry

C ( t ,τ ) = C0 ª¬1 − e ( ) º¼ + C1e − γ 2τ ª¬1 − e 3 ( ) º¼ ,


− γ t −τ − γ t −τ
(6)

where the differential equation will have a second order. Measure the creep of
the Alexander S. lead to differential equations of the fifth order. Pay attention to the
important requirement to the function C(t,IJ). The equations of second and higher
orders appear drag force proportional to the total acceleration ε(t ) . This is
incompatible with Newtonian mechanics and violates its fundamental principle of the
independence of the forces.

Saving the form action creep Harutyunyan N., Prokopovich I., Karapetyan K.,
we obtain the differential equation of the first order that is convenient for practical
tasks

a ( t ) ε M ( t ) + δ ( t ) ε M2 ( t )
ε ( t ) + γε ( t ) = εM ( t ) + γθ ( t ) + γε M ( t ) , (7)
1 + g (t ) ε M (t )

or

© ASCE
CONCREEP 10 244

Esec ª¬σ ( t ) º¼
Esec ª¬σ ( t ) º¼ ª¬ε ( t ) + γε ( t ) º¼ = σ ( t ) + γ {1 + θ ( t ) Esec ª¬σ ( t ) º¼}σ ( t ) . (8)
Etan ª¬σ ( t ) º¼

Let’s pay attention to the important results. The presence in the creep law (7)
instantaneous deformation İɦ concrete and its relation to the total strain İ opens
opportunities for almost an exact solution of a number of fundamental design
problems, such as long-term resistance to compression of concrete columns and
beams ( Sanzharovsky R., 1978). It is possible, under the conditions of nonlinear
creep and instantaneous nonlinearity of concrete, to take into account the exact stress
distribution over the cross section within the deformation model of Eurocode 2, and
thereby bridge the gap that exists between the theories of short-term and prolonged
resistance reinforced concrete structures. It is possible in the evaluation of long-term
ultimate bearing capacity of structures to apply the criterion to achieve the
normalized value of the limiting compressive deformation İɜ2 Eurocode 2. Parallel
accounted for the change over time of concrete strength RB.

3. EXAMPLE

For example, consider the definition of the carrying capacity of eccentrically


compressed main regulatory elements in the design scheme. After integration over
the cross section on the basis of deformation model section of Eurocode 2 and
diagrams (5) we can write the equilibrium condition:

­ ba c
° N = g 2 ε ª¬ gε 2 ɦ − ln (1 + gε 2 M ) º¼ +
° 2M

° bδ c ª 1 2 2 º
°+ 3 « g ε 2 M − gε 2 M + ln (1 + gε 2 M ) » +
° g ε 2M ¬ 2 ¼
° ' ' ε2 ε2
° + As Es ( c − a ) − As Es ( h0 − c ) ,
'

° c c
® (9)
° M = N ( e + f ) = bδ c ª 1 g 3ε 3 − 1 g 2ε 2 + gε − ln (1 + gε ) º +
2

g 4 ε 22M «¬ 3
2M »
° 2M
2
2M 2M
¼
°
° + bδ c ª 1 g 2ε 2 − gε + ln 1 + gε ' ' ε2
2
º
( c − a' ) +
2

° g 3 ε 2 «¬ 2 2M 2M ( 2 M ) » − As Es
¼ c
° 2M

° ' ' ε2
°̄+ As Es c ( h0 − c ) ,
2

where b, h0 - cross-sectional dimensions of the column; c - the height of the


compression zone of the section; h0-c - depth of cracks in the concrete; İ2, İ2ɦ -
complete and instantaneous deformation extreme fiber compressed zone; a, į, g -
regulatory options diagram by Sarzhin; e - eccentricity; f - the deflection of the
middle column.

© ASCE
CONCREEP 10 245

Equilibrium equation is differentiable with respect to t

­ ∂N  ∂N ∂N
° ∂ε ε 2 ( t ) + ∂c c ( t ) + ∂ε ε2 M ( t ) = 0,
° 2 2M
® (10)
°N ∂f ∂ f ∂M ∂M ∂M
ε2 ( t ) + N c ( t ) = ε2 ( t ) + c ( t ) + ε2 M ( t ) = 0.
°¯ ∂ε 2 ∂c ∂ε 2 ∂c ∂ε 2 M

If using a second-order equation of creep, the equation (10) is differentiated


once again and so on.
From equations (10), read in conjunction with (7), we have a complete system
of governing equations for calculating the deformation of reinforced concrete
columns under creep conditions, which can easily transform (mean linearity) to the
normal form of Cauchy

ε2 ( t ) = Φ 2 ª¬ε 2 ( t ) , ε 2 M ( t ) , c ( t ) º¼ ,
ε2 M ( t ) = Φ 2 M ª¬ε 2 ( t ) , ε 2 M ( t ) , c ( t ) º¼ , (11)
c ( t ) = Φ c ª¬ε 2 ( t ) , ε 2 M ( t ) , c ( t ) º¼ .

Integration of (11) is carried out according to standard programs; it repeatedly


performed as the authors themselves and their numerous students. Settlement of (11)
is accompanied by three independent test conditions characterizing achievement
column limit state, and indicates a loss of bearing capacity:

a) ε 2 M ( t ) ≤ ε B 2 = 350 ⋅10 −5 in the compressed zone of concrete


h 0 −c ( t ) Rs
b) ε1s = ε 2 ( t ) ≤ in the tension reinforcement
c (t ) Es
c) the condition of the critical state, based on the calculation of the second-
order nonlinear.

Condition critical condition characterized by the loss of stability of the


column moment during creep, according to the data, is found by varying the
isochronous quasistatic equilibrium equations:

­ ∂N ∂N ∂N
° ∂ε δε 2 + ∂c δ c + ∂ε δε 2 M = 0,
° 2 2M

° ∂f ∂f ∂M ∂M ∂M
®N δε 2 + N δ c = δε 2 + δc + δε 2 M = 0, (12)
° ∂ε 2 ∂c ∂ε 2 ∂c ∂ε 2 ɦ
°δε 2 = δε 2 M .
°
¯

© ASCE
CONCREEP 10 246

Determinant of the system (12) is equal to zero, is the condition of the critical
state. This determinant is calculated together with the system of equations (11). It
captures the moment of the ultimate state, the time loss of bearing capacity of the
structure, including the duration of the existence of the structure.

4. CONCLUSIONS

Developed theory of partitions and covers short-term bearing capacity of the


structure by identifying the values of İ2, İ2M. This theory covers cases of alternating
periods of short-term and long-term existence of reinforced concrete structures. In
this theory, a unified manner is taken into account together six types of nonlinear
behavior of reinforced concrete structures:
1. Instantaneous deformation elastoplastic concrete;
2. The elastic-plastic deformation of the valve;
3. The presence of cracks in the tension zone of the concrete;
4. The linear and nonlinear creep of concrete;
5. Creep special types of reinforcement (reinforced plastics, polymers, etc.),
which leads to higher order differential equations;
6. The two types of geometric nonlinearity in the approximate and exact
expression for the curvature of the curved axis of the column.
According to the results of the theory developed practical methods of
calculation of reinforced concrete structures, which are more reliable and cost-
effective for the following reasons. The theory is based on estimated Principles and
Rules of Eurocode 2 corresponds to the current state of the theory of elastic-plastic
analysis of structures and foundations of classical Newtonian mechanics. In theory
not used questionable or controversial assumptions. Computational procedures are
based on the classical method of numerical solution of the Cauchy problem for
normal systems of differential equations. Numerical results are consistent with the
theory of multiple reliable experiments Tal K. and Chistyakov E. conducted in the
Research Institute of Concrete. They also coincide with the limit values obtained
during the transition to the linear theory. Used functions and their parameters
describing the phenomenological creep curves of concrete substantiated by numerous
experiments of famous scientists.
We can show the results for practical design in the table, which lists the
expansion coefficient of resistance reinforced concrete structure. They demonstrate,
in accordance with the standard method, to increase economic efficiency for ordinary
consumers, as well as economy of materials and energy costs at the State level.
Table 1. is an example and contains a set of coefficients prolonged resistance
compressed concrete structures corresponding ij’=0,8. It is compared with a standard
value 0,555 and demonstrates the great convention of the existing standard for
concrete in terms of accounting creep of concrete.

© ASCE
CONCREEP 10 247

Table 1. Prolonged resistance coefficients


Value Eccentricity
reinforcement 0,1 0,2 0,3 0,4 0,5 0,8
0,2 0,838 0,845 0,855 0,860 0,875 0,890
0,15 0,830 0,835 0,843 0,850 0,862 0,875
0,125 0,821 0,825 0,828 0,835 0,845 0,855
0,10 0,817 0,820 0,821 0,822 0,828 0,828
0,075 0,815 0,815 0,815 0,815 0,815 0,815
0,05 0,795 0,795 0,791 0,790 0,787 0,785
0,025 0,785 0,780 0,772 0,767 0,755 0,740
0,01 0,765 0,757 0,750 0,737 0,715 0,695
1,0 1,2 1,4 1,6 1,8 2,0
0,2 0,910 0,928 0,945 0,968 0,980 0,990
0,15 0,885 0,900 0,915 0,925 0,937 0,950
0,125 0,865 0,880 0,890 0,900 0,910 0,922
0,10 0,847 0,855 0,865 0,875 0,882 0,895
0,075 0,815 0,815 0,820 0,822 0,825 0,830
0,05 0,782 0,780 0,777 0,775 0,770 0,765
0,025 0,730 0,712 0,690 0,670 0,650 0,625
0,01 0,670 0,645 0,625 0,600 0,575 0,550

REFERENCES
Sanzharovsky, RS 1978, Stability of construction elements during creep,
LSU. (in Russian)
Sanzharovsky, RS 2013, 'Problems of the theory of creep', Structural
Mechanics engineering structures and buildings, no. 3, pp. 28-34. (in Russian)
Sanzharovsky, RS 2014, 'Nonlinear hereditary creep theory', Building
mechanics engineering structures and buildings, no. 1, pp. 63-68. (in Russian)
.

© ASCE
CONCREEP 10 248

Massive Structure Monitoring: Relevance of


Surface Strain Measurement

Maxime Boucher1; Matthieu Briffaut1; and Frédéric Dufour1,2


1
Univ. Grenoble Alpes, 3SR, F-38000 Grenoble, France; and CNRS, 3SR, F-38000
Grenoble, France.
2
Chair Professor PERENITI.
E-mail: maxime.boucher@3sr-grenoble.fr

Abstract

Most of large civil engineering concrete structures have been instrumented for
decades with embedded sensors. To prevent the eventual loss of data, complementary
instrumentation of external surface has recently been deployed. This new
instrumentation can take different forms but in all cases, to avoid damaging the
structure, it will be only superficially anchored. Near the outer surfaces, thermo-hydro-
mechanical concrete behaviour is more sensitive to varying environmental conditions
than in the centre of the structures. Therefore, the strain measured near the outer surfaces
is not identical to the strain measured by embedded sensors. Consequently the methods
of classical physical-statistical analysis must be reviewed. Using a thermo-hydro-
mechanical finite element modeling calibrated on a representative concrete and applied
on a current part of a thick structure, this work confirms a dependence of strain on the
depth. First results show that the depth impact affects both kinetic and amplitude strain.

Keywords:
Thermo-hydro-mechanical behavior; Concrete; Massive structure; Variable
environnemental conditions; Surface monitoring.

INTRODUCTION

For nearly 80 years, most of large civil engineering concrete structures are
instrumented with vibratory strain gauge. On some structures such as confinement
vessels of nuclear power plants (highly prestressed), operators have to cope with the

© ASCE
CONCREEP 10 249

loss of signal from some sensors. This can be explained by several reasons including a
default in the electrical system that excite the wire or by the excess of frequency limit
measurement due for instance to important creep strains (Simon and Courtois, 2011;
Simon et al. 2013). Most of these sensors have been embedded in the concrete during
construction and are thus irreplaceable. To overcome this loss of information, one of
the solutions is to replace these defective embedded sensors by surface sensors
(Figure 1).

Figure
1:
Embedd
ed strain
sensor
(a) and
surface
strain
sensor
(b).

Nevertheless, unlike central strain (far from outer surface), skin strain (near outer
surfaces) is strongly affected by the environmental condition variations. Temperature
and humidity cycles (daily and seasonal) generate thermo-hydric variations of volume.
Other phenomena such as sunshine, rain or wind, also disrupt these two main fields.
Finally, characteristic times and so influence depths of these phenomena are very
different. For these reasons, the analysis of this new data is more complex than
embedded measures for which usual statistical analysis methods exist.
The aim of the present contribution is to compare the strain evolution through the
depth in thick structure. This can be considered as a first step toward the complete
definition of a space and time transfer functions between both strain vs time signals.

For this purpose, a visco-elastic finite element model able to represent thermo-hydro-
mechanical behaviour of concrete submitted to variable thermo-hydric boundary
conditions is presented in the first part. Then, the calibration of the material constitutive
parameters on laboratory tests is detailed. Finally, the strains at different depths from a
simulation over thirty years on a current part of containment enclosure submitted to
thermo-hygrometric variations are analysed.

© ASCE
CONCREEP 10 250

NUMERICAL MODEL

Numerical strategy.
To be applicable to major civil engineering thick structures, the numerical model must
account for thermal and hydric variation effects on strains. Thermo-mechanical
(Arthanari and Yu, 1967 ; Seki and Kawasumi, 1972 ; Kommendant et al, 1976),
hydro-mechanical (Wittmann, 1970 ; Wittmann, 1973) and thermo-hydric (Caré,
2008) chaining are retained. In contrast, the effects of mechanical strains on thermal
and hydric properties (Lassabatere et al, 1997) and the modification of heat and water
diffusivity due to cracked skin are neglected. Finally, for ageing structures, hydration
reaction is supposed finished.

Complete simulations are carried out using Code_Aster by chaining three calculations:
thermal field, hydric field and mechanical strains. Constitutive models adopted are for
the most part empirical although based on physical approaches.

Thermal field.
Thermal field is assumed to be governed by a conventional linear heat equation with
convective boundary conditions (Neumann type):

{
ȡ Cth TÚ = k th ǻT
ࢥ th = hth( T í T ’ )
(1)

Specific heat capacity ȡCth, thermal conductivity kth and heat transfer coefficient hth
(taking into account convection and radiation) are taken constant during all the
calculation.

Hydric field.
Drying process is modeled by a nonlinear diffusion equation involving a diffusion
coefficient (non-linear function of water content) usually written as:
CÚ = 䳱[ D(C , T ) 䳱T ] (2)
Several authors have proposed various relationships between drying coefficient and
water content. The expression used, proposed by Granger (1995), is composed of the
drying coefficient proposed by Mensi (1988) and the temperature dependence given
by Bazant (1972):
íQ 1 1
( í )
bC T R T T0
D(C , T ) = a e e (3)
T0

© ASCE
CONCREEP 10 251

The coefficients a,b and the activation energy Q of Arrhenius' law are supposed to be
constant. T0 represents the reference temperature.
The calculation of drying is done by means of water content, although boundary
conditions are prescribed in relative humidity. The isotherm between relative humidity
of concrete and water content must be entered. In this study, the hysteresis existing
between sorption and desorption cycles is neglected and the isotherm is supposed
linear in the humidity range between 40% and 100%. Therefore, the used isotherm
function is:
C eq [h 0í h] í C0 [ hí heq ]
C( h) = (4)
h0í heq

where h0, C0 are the relative humidity and initial moisture content and heq, Ceq are the
relative humidity and water content at equilibrium.
The convective boundary condition used involves a water exchange coefficient hdr
which is assumed constant:
ࢥ dr = hdr [ Cí C( hair )] (5)
By analogy with nonlinear thermal calculation, water field is obtained using the non-
linear thermal module of Code_Aster.

Mechanical field.
This study focuses on operating structures during service life. Consequently damage
rate is assumed to be low and a simple elastic model is used.
Under the assumption of small strains, the total strain tensor is assumed to be the
sum of five tensors: elastic strain (İEl), thermal strain (İTh), drying shrinkage strains
(İDS), basic creep strains (İBC) and drying creep strains (İDC):
İtot = İEl + İTh + İDS + İ BC + İ DC (6)

• The thermal strain is assumed to be proportionnal to the temperature change:


İEl = Į Th . TÚ .1 .
• For the calculation of drying shrinkage, a phenomenological model assuming
Ú
proportionality with water variation content is used: İDS = Į DS . C . 1 .
• The basic creep model is based on simple rheological models: an elastic body, a
linear Kelvin-Voigt for modeling reversible creep and a Maxwell unit with a
nonlinear viscosity to model the long time creep. The model also assumes
complete decoupling between the spherical and deviatoric components. Both
of these chains are however equivalent in their construction. A linear
dependance with internal relative humidity is included.

© ASCE
CONCREEP 10 252

• The model, proposed by Bazant (1985), is assumed to be a linear relation with


İDC = Į DC . hÚ . ı
strain and internal relative humidity variations: .

CALIBRATION

The numerical model presents 20 material parameters to be determined. For


this, laboratory tests on specimens formed by a representative concrete of nuclear
power plant are used. These tests are of two types:
• A first test campaign on 4 specimens of dimensions 25x50cm subjected to axial
compression cycles from 0 to 12 MPa.
• A second test campaign on 14 specimens of dimensions 16x100cm subjected to
various conditions:
• Non-drying weight loss (1 specimen)
• Drying weight loss (1 specimen)
• Non drying shrinkage (3 specimens)
• Drying shrinkage (3 specimens)
• Non drying creep (3 specimens)
• Drying creep (3 specimens)
Non-drying conditions were ensured by strips of self-adhesive aluminium. The
temperature was maintained near to 20°C. The relative humidity has oscillated
between 50% and 70%. All strain measurements were carried out with inductive
sensors located outside the specimens.
Since the variability between specimen responses from the second campaign was low,
only one out of the three (for shrinkage and creep tests) was carried out up to the end
of the campaign. The model was calibrated with these last specimens.
For confidentiality reasons, the experimental results will be normalized and the
parameter values will not be provided.

Thermal field.
The thermal diffusion and the specific capacity cannot be measured out of these tests.
Thus, their retained values were obtained from the literature. It is the same for the
heat exchange coefficient.

Hydric field.
Before calibrating the drying model, the sorption-desoprtion isotherm of the concrete
must be determined in order to translate the boundary conditions (known in relative

© ASCE
CONCREEP 10 253

humidity of the surrounding air) in concrete water content.


Initial relative humidity of concrete (h0) is taken equal to 100% (saturated) and equal
to the average of the relative humidity in the air during all the test at equilibrium
(heq).
The initial water content (C0) is defined as the amount of water used to make
concrete minus the water used in hydration reaction (Granger 1995). The final water
content (Ceq) is obtained by extrapolating the weight loss curve to infinity.
Finally, the drying model was calibrated on the drying mass loss test. The water
exchange coefficient and the coefficient b (Equation 3) from diffusion coefficient have
been fixed to reference values from the literature. Only the coefficient a from diffusion
coefficient has been calibrated. The comparison between the experimental points and
the numerical curve is presented in Figure 2a.

Elastic strain.
Measurements from the first campaign were used to determine the Young's modulus of
concrete studied. A linear regression of the stress-strain curves (Figure 3a) allows to
determine an average Young's modulus for these four specimens.
Since only axial strains were measured, the Poisson's ratio could not been
determined. Therefore a standard value for concrete is used.

Thermal strain.
No thermal expansion test was made. But peaks of temperature up to 25°C were
measured during the campaign. These variations have been felt on shrinkage and
creep tests. These thermal anomalies have been used to calibrate the thermal
expansion coefficient.
The analysis of the experimental results yields an increase of strain with temperature
amplitude (Figure 3b). A linear regression of these points gives direct access to the
thermal expansion coefficient ĮTh.

© ASCE
CONCREEP 10 254

Figure 2: Calibration of the thermo-hydro-mechanical model. Numerical simulations:


weight loss (a) and total creep (b).

Drying shrinkage.
The drying shrinkage strain is assumed to be proportional to the variation of water
content. Therefore, only one parameter must be identified (ĮDS). Autogenous
shrinkage was estimated at short-term from non-drying weight loss. Then, it was
substracted to drying shrinkage test in order to obtain the drying shrinkage. The
obtained curve is not exactly linear at the beginning (Figure 3c). Indeed, at the start
of drying, the high tensile stress located near the exchange surfaces creates skin
cracking, which relaxes the stress.
A linear regression of the central portion of the curve allows the calibration of the
drying shrinkage parameter. Numerically, the strain obtained will be greater than the
experimental measurement. However both kinetic and amplitude related to relative
humidity cycles will be respected.

Drying creep.
It was decided arbitrarily to calibrate drying creep before basic creep (since it has
only one parameter). Experimental curve of drying creep was obtained by subtracting
non-drying creep and drying shrinkage to the drying creep test. Since drying creep
was assumed proportional to the stress and the relative humidity variation, drying
creep values were divided by the applied stress. Meanwhile, the mass loss was
translated into relative humidity thanks to the sorption-desorption isotherm.
A quasi-linear increase of drying creep with relative humidity variation is highlighted
by the analysis of experimental results in Figure 3d. Again, as for drying shrinkage,
the resulting curve is not perfectly linear in its ends. For the beginning, explanations
may be the same than in the case of drying shrinkage. For the long term, the

© ASCE
CONCREEP 10 255

permeability of dessiccation protection is probably involved. It was found that after a


few years, the mass loss from non-drying test increased significantly. Consequently,
in our approach, by subtracting non-drying creep and drying shrinkage to drying
creep, drying shrinkage is removed twice which explains the decrease at the end of the
curve.
That is why, drying creep parameter is identified by linear regression of the central
portion of the experimental curve only (Figure 3d).

Figure 3: Calibration of the thermo-hydro-mechanical model. Linear regressions:


elastic strain (a), thermal strain (b), drying shrinkage (c) and drying creep (e).

Basic creep.
Finally, basic creep is calibrated on the total creep test, taking into account the
previously calibrated drying creep. Experimental and numerical results are presented in
Figure 2b.

SIMULATION METHOD

After a thorough calibration of material parameters for each modeled


phenomena, an application to a representative volume of a confinement vessel of a
nuclear power plant was achieved. For this simulation, an inner containment wall of a

© ASCE
CONCREEP 10 256

double containment vessel is studied which allows the suppression of the direct effects
of rain, wind and sunshine; only thermal and humidity variations, load and prestressing
were selected as thermo-hydro-mechanical boundary conditions.
The mesh of a current portion of containment in three dimensions (width of 2 meters,
height of 1.8 meters, thickness of 1.2 meters and internal radius of 21.9 meters),
taking into account reinforcements was used (Figure 4).

Figure 4: Geometry and dimensions of the studied structure.

The calculation simulates 30 years of service life of the structure. The thermo-
hygrometric boundary conditions were created by duplicating periodic temperature
and humidity measurements from the studied structure.

In-situ mechanical boundary conditions are represented by the following hypothesis:


on the bottom and lateral faces, normal displacements are prohibited and the upper
side is forced to remain parallel to the bottom side.

During the first three years, only the structural loading, the thermal and the moisture
variations (identical on the inner and outer faces) are taken into account.
At 3.5 years, the vertical and horizontal preloading is added.
At 6 years, the reactor is operating. This yields a change in thermo-hygrometric
boundary conditions (different on the inner and outer sides).
At 16 and 26 years, two containment pressure tests are simulated (pressurizing the
reactor vessel for 3 days). For that, an overpressure is added on the inner side and
an depression is added on the upper face (to represent the vertical stretch of the
structure).

© ASCE
CONCREEP 10 257

RESULTS :

The orthoradial strains obtained numerically on the outer surface and near the
center of the structure are presented in Figure 5. Also for confidentiality reasons, the

strains are normalized by the absolute value of the center strain at 30 years.
Figure 5: Tangential strain on outer surface and close to center of structure.

On this graph, one can remark that:


• Strain cycles related to thermal variations between the center and the skin are
closed to each others.
• Given the thickness of the structure and a very low water diffusivity, in
comparison to thermal strain and creep, the drying shrinkage of the concrete is
only slightly perturbed by the changes in humidity of the air.
• Before prestressing, both strains observed in skin and in the center are
similar. After prestressing the cables, the creep kinetics begin to dissociate:
faster in the center (high water content) than at the surface (low water
content). After 30 years, this difference is about 5%.

CONCLUSION

The results of this study show that, unlike thermal variations which affect the
entire thickness of the structure, creep kinetics are not the same between the center and
the surface of the structure. Creep is smaller near the exchange surfaces, where

© ASCE
CONCREEP 10 258

concrete has a lower water content.


Under the hypothesis used in the numerical part, this contribution shows that the
strain evolution between the core and the surface of the structure are not exactly the
same. This confirms that the identification of a transfer function is necessary between
strain measurement at various depth. Given the above results, a simple reduction of
the strain depending on the distance to the outer sides could be accurate enough.
Benefiting from a representative structure, built in the same concrete used for the
calibration procedure and instrumented with surface strain sensors for several years,
this present work will continue with a comparison between simulated strain and
measured strain.
Acknowledgments.

Work performed thanks to the support of EDF in context of Chair PERENITI run by
the Fondation Partenariale Grenoble INP and using the monitoring data provided by
EDF-DTG. Partners responsibility of the Chair cannot in any circumstances be
blamed on the grounds of the content of the publication, which is only binding its
author.

REFERENCES

Arthanari S., Yu C.W. (1967). “Creep of conrete under uniaxial and biaxial stresses
at elevated temperatures”. Mag. Concrete Res. 13(60), 149-156.
Bazant Z.P., Najjar L.J. (1972). “Nonlinear water diffusion in nonsaturated concrete”.
Materials and Constructions, 1972, Vol. 5, 3-20.
Bazant Z.P., Chern J.C. (1985). “Concrete creep at variable humidity: constitutive
law and mechanism”. Matériaux et Constructions, Vol. 18, n°103, pp 1-20.
Caré S. (2008). “Effect of temperature on porosity and on chloride diffusion in cement
pastes”. Construction and Building Materials 22, 1560-1573.
Code_Aster, Électricité De France (EDF), Code_Aster finite element code version
11.5, available at http://www.code-aster.org.
Granger L. (1995). “Comportement différé du béton dans les enceintes de centrales
nucléaires: analyse et modélisation”. Laboratoire Central des Ponts et
Chaussées.
Kommendant G., Polivka M., Pirtz D. (1976). “Study of concrete properties for
prestressed concrete reactor vessels”. Rapport technique, Department of Civil
Engineering.
Lassabatere T., Torrenti J.-M., Granger L. (1997). “Sur le couplage entre séchage du

© ASCE
CONCREEP 10 259

béton et contrainte appliquée”. Pages 331-338.


Mensi R., Acker P., Attolou A. (1988). “Séchage du béton : analyse et modélisation”.
Materials and Structures/Matériaux et Construction, 1988, 21, 3-12.
Seki S., Kawasumi M. (1972). “Creep of concrete at elevated temperatures”.
Concrete of Nuclear Reactors, 1:591-638.
Simon A., Courtois A. (2011). « Structural Monitoring of Prestressed Concrete
Containments of Nuclear Power Plants for ageing management ». Transactions,
SmiRT 21, 6-11 November, 2011, New Delhi, India.
Simon A., Oukhemanou E., Courtois A. (2013). « Structural Monitoring of
Prestressed Concrete Containments of Nuclear Power Plants for ageing
management ». Technical Innovation in Nuclear Civil Engineering – TINCE
2013, Paris (France), October 23-31, 2013.
Wittmann F. (1970). « Einfluss des feuchtigkeitsgehaltes auf das kriechen des
zementsteines ». rheologica Acta, 9(2):282-287.
Wittmann F. (1973). « Interaction of hardened cement paste and water ». Journal of
the American ceramic society, 56(8):409-415.

© ASCE

Você também pode gostar